Solving the Impossible in Quantum Field Theory

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments

This is Awesome. The last few episodes building up to this one, along with the promised ramp up into the Standard Model, really makes this channel feel like a curriculum. I feel like I'm actually getting educated when I watch. I love it!

👍︎︎ 7 👤︎︎ u/PapaTua 📅︎︎ Jul 13 2017 🗫︎ replies
Captions
[MUSIC PLAYING] HOST: This episode is supported by Curiosity Stream. Quantum field theory is stunningly successful at describing the smallest scales of reality, but its equations are also stunningly complex. A lot of the genius in QFT's development was in finding brilliant hacks to make these equations workable. The most famous of these are the incredible Feynman diagrams. [MUSIC PLAYING] The equations of quantum field theory allow us to calculate the behavior of subatomic particles by expressing them as vibrations in quantum fields. But even the most elegant and complete formulations of quantum field theory, like the Dirac equation or Feynman's path integral, become impossibly complicated when we try to use them on anything but the most simple systems. But physicists tend to interpret "that's impossible" as "I dare you to try," and try they did. First, they expressed these impossible equations in approximate, but solvable, forms. Then they tackled the pesky infinities that kept appearing in these new approximate equations. Finally, the entire mess was ordered into a system that mere humans could deal with using the famous Feynman diagrams. To give you an idea of how messy quantum field theory can be, let's look at what should be a simple phenomenon-- electron scattering, when two electrons repel each other. In old-fashioned classical electrodynamics, we think of each electron as producing an electromagnetic field. That field then exerts a repulsive force on the other electron. At least in the simplest cases, the Coulomb equation governing this subatomic billiards shot is really easy to solve. But in quantum field theory, specifically quantum electrodynamics, or QED, the story is very different. We think of the electromagnetic field as existing everywhere in space, whether or not there's an electron present. Vibrations in the EM field are called photons, what we experience as light. The electron itself is just an excitation, a vibration in a different field-- the electron field. And the electron and EM fields are connected. Vibrations in one can cause vibrations in the other. This is how QED describes electron scattering. One electron excites a photon, and that photon delivers a bit of the first electron's momentum to the second electron. It's arguable exactly how real that exchanged photon is. In fact, we call it a virtual photon, and it only exists long enough to communicate this force. There are other types of virtual particle whose existence is similarly ambiguous. We'll get back to those in another episode. This is a good time to introduce our first Feynman diagram. The brilliant Richard Feynman developed these pictorial tools to organize the painful mathematics of quantum field theory, but they also serve to give a general idea of what these interactions look like. In a Feynman diagram, one direction is the time-- in this case, up. The other axis represents space, although the actual distances aren't relevant. Here we see two electrons entering in the beginning and moving towards each other. They exchange a virtual photon-- this squiggly line here-- and the two electrons move apart at the end. But Feynman diagrams aren't really just drawings of the interaction. They're actually equations in disguise. Each part of the Feynman diagrams represents a chunk of the math. Incoming lines are associated with the initial electron states, and outgoing lines represent the final electron states. The squiggle represents the quantized fueled excitation of the photon, and the connecting points, the vertices, represent the absorption and emission of the photon. The equation you string together from this one diagram represents all of the ways that two electrons can deflect involving only a single virtual photon. And from that equation, it's possible to perfectly calculate the effect of that simple exchange. Unfortunately, real electron scattering at a quantum level is a good deal more complicated than this. For that reason, this simple calculation gives the wrong repulsive effect between two electrons. If we observe two electrons bouncing off each other, all we really see is two electrons going in and two electrons going out. The quantum event around the scattering is a mystery. There are literally infinite ways that scattering could have occurred. In fact, according to some interpretations, all infinite intermediate events that lead to the same final result actually do happen, sort of. We talked about this weirdness when we discussed the Feynman path integral recently. Just as with the path integral, to perfectly calculate the scattering of two electrons, we need to add up all of the ways the electrons can be scattered. And this is where Feynman diagrams start to come in handy, because they keep track of the different families of possibilities. For example, the electrons might exchange just a single virtual photon, but they might also exchange two, or three, or more. The electrons might also emit and reabsorb a virtual photon. Or any of those photons might do something crazy, like momentarily split into a virtual anti-particle-particle pair. Those last two events are actually hugely complicating, as we'll see. With infinite possible interactions behind this one simple process, a perfectly complete quantum field theoretic solution is impossible. But if you can't do something perfectly, maybe near enough is good enough. This is the philosophy behind perturbation theory, an absolutely essential tool to solving quantum field theory problems. The idea is that if the correct equation is unsolvable, just find a similar equation that you can solve, then make small modifications to it-- perturb it-- so it's a bit closer to the equation that you want. It'll never be exact, but it might get you pretty close. In the case of electron scattering, the most likely interaction is the exchange of a single photon. Every other way to scatter the electrons contributes less to the probability of the event. In fact, the more complicated the interaction, the less it contributes. Here, Feynman diagrams are indispensable. It turns out that the probability amplitude of a particular interaction depends on the number of connections, or vertices, in the diagram. Every additional vertex in an interaction reduces its contribution to the probability by a factor of around 100. So the most probable interaction for electron scattering is the simple case of one photon exchange with its two vertices. A three-vertex interaction would contribute about 1% of the probability of the main two-vertex interaction. However, it turns out that for electron scattering, there are no three-vertex interactions. However, there are several interactions that include four vertices, and each contributes about 1% of 1% of the two-vertex interaction, and this is true even though those complex interactions are very different to each other. They include exchanging two virtual photons, or one electron emitting and reabsorbing a virtual photon, or the exchanged photon momentarily exciting a virtual electron-positron pair. And more complicated interactions add even less to the probability. So with Feynman diagrams, you very quickly get an idea of which are the important additions to your equation and which you can ignore. Perturbation theory, with the help of Feynman diagrams, make the calculation possible, but that doesn't mean we're done. Including all of these weird intermediate states really opens up a can of worms. This is especially true for so-called loop interactions, like when a photon momentarily becomes a virtual particle-anti-particle pair and then reverts to a photon again, or when a single electron emits and reabsorbs the same photon. This latter case can be thought of as the electron causing a constant disturbance in EM field. Electrons are constantly interacting with virtual photons. This impedes the electron's motion and actually increases its effective mass. The effect is called self-energy. But if you try to calculate the self-energy correction to an electron's mass using quantum electrodynamics, you get that the electron has infinite extra mass. This sounds like a problem. To calculate the mass correction due to one of these self-energy loops, you need to add up all possible photon energies, but those energies can be arbitrarily large, sending the self-energy-- and hence, the mass-- to infinity. In reality, something must limit the maximum energy of these photons. We don't know what that limit is. The answer probably lies within a theory of quantum gravity which we don't yet have. But just as with perturbation theory, physicists found a cunning trick to get around this mathematical inconvenience. It's called renormalization. Obviously, electrons do not have infinite mass, and we know that because we've measured that mass, although any measurement we make actually includes some of this self-energy, so our measurements are never of the fundamental or bare mass of the electron, and that is where the trick lies. Instead of trying to start with the unmeasurable fundamental mass of the electron and solve the equations from there, you fold in a term for the self-energy corrected mass based on your measurement. In a sense, you capture the theoretical infinite terms within an experimental finite number. This renormalization trick can be used to eliminate many of the infinities that arise in quantum field theory-- for example, the infinite shielding of electric charge due to virtual particle-anti-particle pairs popping into and out of existence. However, you pay a price for renormalization. For every infinity you want to get rid of, you have to measure some property in the lab. That means the theory can't predict that particular property from scratch. It can only make predictions of other properties relative to your lab measurements. Nonetheless, renormalization saved quantum field theory from this plague of infinities. Feynman diagrams successfully describe everything from particle scattering, self-energy interactions, matter-anti-media creation and annihilation, to all sorts of decay processes. We'll go further into the nuts and bolts of Feynman diagrams in an upcoming challenge episode. A set of relatively straightforward rules governs what diagrams are possible, and these rules make Feynman's doodles an incredibly powerful tool for using quantum field theory to predict the behavior of the subatomic world. The results led to the standard model of particle physics. In future episodes, we'll talk more about what is now the most complete description we have for the smaller scales of space time. This episode is brought to you by Curiosity Stream, a subscription streaming service that offers documentaries and nonfiction titles from some of the world's best filmmakers, including exclusive originals. The two-episode documentary "The Ultimate Formula" gives a really nice history of the development of quantum field theory. Get unlimited access today, and for our audience, the first two months are free if you sign up at curiositystream.com/spacetime and use the promo code spacetime during the sign-up process. As always, a huge thanks to our supporters on Patreon. This week, we'd like to give an extra big thanks to Eugene Lawson, who's contributing at the Big Bang level. Eugene, it's an incredible help, and you make our jobs much easier. And another reminder to current or would-be Patreon patrons-- we made some "Space Time" eclipse glasses-- super handy for not going blind watching the great American eclipse in August. We're sending out a set to every Patreon contributor at the $5 level or above, and that includes anyone who signs up at or increases to the $5 level in the month of July, at least until we run out of glasses. It'll be first come, first served. Last week, we talked about Richard Feynman's brilliant contribution to the development of quantum field theory with his path integral formulation. You guys had a lot to say. Christian Haas asked how Feynman's path integral method, which is compatible with special relativity, can derive Schrodinger's equation when Schrodinger's equation is not compatible. So the deal is that Schrodinger's equation is a special case of a more general formulation of quantum mechanics. In Schrodinger's equation, all of the particles are tracked according to one universal master clock. In Feynman's approach, each particle is tracked according to its own proper time clock, which can vary in its tick speed depending on how fast the particle is traveling. Derivation of Schrodinger from Feynman requires approximating all of the separate proper time coordinates to give a single time coordinate. That approximation is OK at low speeds but breaks when things get close to the speed of light. [INAUDIBLE] points out that the final probability for a particle journey is the square of the length of the complex probability amplitude vector. And yeah, that's right. Probability is the square of the probability amplitude. That's the Born rule right there. However, the sense I wanted to relay is that the total probability depends on the length of the summed probability amplitudes-- so the square of the real plus the square of the complex components. [INAUDIBLE] also correctly points out that the individual paths don't have different probability amplitude lengths taken separately, but rather, they're pointing in the complex vector space, rotates so that each path adds differently to the total probability. [INAUDIBLE] points out that the wildly divergent paths would require superluminal speeds to reach their destination at the same time as the straight line paths. Yeah, that's right. Feynman didn't limit particle velocity to the speed of light. By applying the principle of least action in the determination of probability amplitudes, it turns out that the "crazy" paths, including the superluminal ones, cancel out and add little to the probability.
Info
Channel: PBS Space Time
Views: 949,415
Rating: undefined out of 5
Keywords: pbs, education, space, time, physics, astrophyscs, particle physics, quantum, mechanics, quantum field theory, field, theory, feynman, richard feynman, dirac, schrodinger, infinite, infinity, path, path integral, electrodynamics, electron, interaction, photon
Id: oQ1WZ-eJW8Y
Channel Id: undefined
Length: 15min 21sec (921 seconds)
Published: Wed Jul 12 2017
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.